Pendulum (mechanics)

Last updated
Animation of a pendulum showing the velocity and acceleration vectors. Oscillating pendulum.gif
Animation of a pendulum showing the velocity and acceleration vectors.

A pendulum is a body suspended from a fixed support such that it freely swings back and forth under the influence of gravity. When a pendulum is displaced sideways from its resting, equilibrium position, it is subject to a restoring force due to gravity that will accelerate it back towards the equilibrium position. When released, the restoring force acting on the pendulum's mass causes it to oscillate about the equilibrium position, swinging it back and forth. The mathematics of pendulums are in general quite complicated. Simplifying assumptions can be made, which in the case of a simple pendulum allow the equations of motion to be solved analytically for small-angle oscillations.

Contents

Simple gravity pendulum

A simple gravity pendulum [1] is an idealized mathematical model of a real pendulum. [2] [3] [4] It is a weight (or bob) on the end of a massless cord suspended from a pivot, without friction. Since in the model there is no frictional energy loss, when given an initial displacement it swings back and forth with a constant amplitude. The model is based on the assumptions:

The differential equation which governs the motion of a simple pendulum is

where g is the magnitude of the gravitational field, is the length of the rod or cord, and θ is the angle from the vertical to the pendulum.

"Force" derivation of ( Eq. 1 )
Figure 1. Force diagram of a simple gravity pendulum. Pendulum gravity.svg
Figure 1. Force diagram of a simple gravity pendulum.

Consider Figure 1 on the right, which shows the forces acting on a simple pendulum. Note that the path of the pendulum sweeps out an arc of a circle. The angle θ is measured in radians, and this is crucial for this formula. The blue arrow is the gravitational force acting on the bob, and the violet arrows are that same force resolved into components parallel and perpendicular to the bob's instantaneous motion. The direction of the bob's instantaneous velocity always points along the red axis, which is considered the tangential axis because its direction is always tangent to the circle. Consider Newton's second law, where F is the sum of forces on the object, m is mass, and a is the acceleration. Newton's equation can be applied to the tangential axis only. This is because only changes in speed are of concern and the bob is forced to stay in a circular path. The short violet arrow represents the component of the gravitational force in the tangential axis, and trigonometry can be used to determine its magnitude. Thus, where g is the acceleration due to gravity near the surface of the earth. The negative sign on the right hand side implies that θ and a always point in opposite directions. This makes sense because when a pendulum swings further to the left, it is expected to accelerate back toward the right.

This linear acceleration a along the red axis can be related to the change in angle θ by the arc length formulas; s is arc length: thus:

"Torque" derivation of ( Eq. 1 )

Equation (1) can be obtained using two definitions for torque.

First start by defining the torque on the pendulum bob using the force due to gravity. where l is the length vector of the pendulum and Fg is the force due to gravity.

For now just consider the magnitude of the torque on the pendulum. where m is the mass of the pendulum, g is the acceleration due to gravity, l is the length of the pendulum, and θ is the angle between the length vector and the force due to gravity.

Next rewrite the angular momentum. Again just consider the magnitude of the angular momentum. and its time derivative

The magnitudes can then be compared using τ = dL/dt

thus: which is the same result as obtained through force analysis.

"Energy" derivation of ( Eq. 1 )
Figure 2. Trigonometry of a simple gravity pendulum. Simple pendulum height.svg
Figure 2. Trigonometry of a simple gravity pendulum.

It can also be obtained via the conservation of mechanical energy principle: any object falling a vertical distance would acquire kinetic energy equal to that which it lost to the fall. In other words, gravitational potential energy is converted into kinetic energy. Change in potential energy is given by

The change in kinetic energy (body started from rest) is given by

Since no energy is lost, the gain in one must be equal to the loss in the other

The change in velocity for a given change in height can be expressed as

Using the arc length formula above, this equation can be rewritten in terms of /dt: where h is the vertical distance the pendulum fell. Look at Figure 2, which presents the trigonometry of a simple pendulum. If the pendulum starts its swing from some initial angle θ0, then y0, the vertical distance from the screw, is given by

Similarly, when y1, then

Then h is the difference of the two

In terms of /dt gives

(Eq. 2)

This equation is known as the first integral of motion, it gives the velocity in terms of the location and includes an integration constant related to the initial displacement (θ0). Next, differentiate by applying the chain rule, with respect to time to get the acceleration

which is the same result as obtained through force analysis.

"Lagrange" derivation of ( Eq. 1 )
Coordinates of a simple gravity pendulum. Coordinates of a simple gravity pendulum.png
Coordinates of a simple gravity pendulum.

Equation 1 can additionally be obtained through Lagrangian Mechanics. More specifically, using the Euler–Lagrange equations (or Lagrange's equations of the second kind) by identifying the Lagrangian of the system (), the constraints () and solving the following system of equations

If the origin of the Cartesian coordinate system is defined as the point of suspension (or simply pivot), then the bob is at

and the velocity of the bob, calculated via differentiating the coordinates with respect to time (using dot notation to indicate the time derivatives)

Thus, the Lagrangian is

The Euler-Lagrange equation (singular as there is only one constraint, ) is thus

Which can then be rearranged to match Equation 1 , obtained through force analysis.

Deriving via Lagrangian Mechanics, while excessive with a single pendulum, is useful for more complicated, chaotic systems, such as a double pendulum.

Small-angle approximation

Small-angle approximation for the sine function: For th [?] 0, the approximation sin th [?] th can be made. Small-angle approximation for sine function.svg
Small-angle approximation for the sine function: For θ ≈ 0, the approximation sin θθ can be made.

The differential equation given above is not easily solved, and there is no solution that can be written in terms of elementary functions. However, adding a restriction to the size of the oscillation's amplitude gives a form whose solution can be easily obtained. If it is assumed that the angle is much less than 1  radian (often cited as less than 0.1 radians, about 6°), or then substituting for sin θ into Eq. 1 using the small-angle approximation, yields the equation for a harmonic oscillator,

The error due to the approximation is of order θ3 (from the Taylor expansion for sin θ).

Let the starting angle be θ0. If it is assumed that the pendulum is released with zero angular velocity, the solution becomes

The motion is simple harmonic motion where θ0 is the amplitude of the oscillation (that is, the maximum angle between the rod of the pendulum and the vertical). The corresponding approximate period of the motion is then

which is known as Christiaan Huygens's law for the period. Note that under the small-angle approximation, the period is independent of the amplitude θ0; this is the property of isochronism that Galileo discovered.

Rule of thumb for pendulum length

gives

If SI units are used (i.e. measure in metres and seconds), and assuming the measurement is taking place on the Earth's surface, then g ≈ 9.81 m/s2, and g/π2 ≈ 1 m/s2 (0.994 is the approximation to 3 decimal places).

Therefore, relatively reasonable approximations for the length and period are: where T0 is the number of seconds between two beats (one beat for each side of the swing), and l is measured in metres.

Arbitrary-amplitude period

Figure 3. Deviation of the "true" period of a pendulum from the small-angle approximation of the period. "True" value was obtained numerically evaluating the elliptic integral. Pendulum period.svg
Figure 3. Deviation of the "true" period of a pendulum from the small-angle approximation of the period. "True" value was obtained numerically evaluating the elliptic integral.
Figure 4. Relative errors using the power series for the period. Pendulum rel error.svg
Figure 4. Relative errors using the power series for the period.
Figure 5. Potential energy and phase portrait of a simple pendulum. Note that the x-axis, being angle, wraps onto itself after every 2p radians. Pendulum phase portrait.svg
Figure 5. Potential energy and phase portrait of a simple pendulum. Note that the x-axis, being angle, wraps onto itself after every 2π radians.

For amplitudes beyond the small angle approximation, one can compute the exact period by first inverting the equation for the angular velocity obtained from the energy method ( Eq. 2 ), and then integrating over one complete cycle, or twice the half-cycle or four times the quarter-cycle which leads to

Note that this integral diverges as θ0 approaches the vertical so that a pendulum with just the right energy to go vertical will never actually get there. (Conversely, a pendulum close to its maximum can take an arbitrarily long time to fall down.)

This integral can be rewritten in terms of elliptic integrals as where F is the incomplete elliptic integral of the first kind defined by

Or more concisely by the substitution expressing θ in terms of u,

 Eq. 3

Here K is the complete elliptic integral of the first kind defined by

For comparison of the approximation to the full solution, consider the period of a pendulum of length 1 m on Earth (g = 9.80665 m/s2) at an initial angle of 10 degrees is The linear approximation gives

The difference between the two values, less than 0.2%, is much less than that caused by the variation of g with geographical location.

From here there are many ways to proceed to calculate the elliptic integral.

Legendre polynomial solution for the elliptic integral

Given Eq. 3 and the Legendre polynomial solution for the elliptic integral: where n!! denotes the double factorial, an exact solution to the period of a simple pendulum is:

Figure 4 shows the relative errors using the power series. T0 is the linear approximation, and T2 to T10 include respectively the terms up to the 2nd to the 10th powers.

Power series solution for the elliptic integral

Another formulation of the above solution can be found if the following Maclaurin series: is used in the Legendre polynomial solution above. The resulting power series is: [5]

more fractions available in the On-Line Encyclopedia of Integer Sequences with OEIS:  A223067 having the numerators and OEIS:  A223068 having the denominators.

Arithmetic-geometric mean solution for elliptic integral

Given Eq. 3 and the arithmetic–geometric mean solution of the elliptic integral: where M(x,y) is the arithmetic-geometric mean of x and y.

This yields an alternative and faster-converging formula for the period: [6] [7] [8]

The first iteration of this algorithm gives

This approximation has the relative error of less than 1% for angles up to 96.11 degrees. [6] Since the expression can be written more concisely as

The second order expansion of reduces to

A second iteration of this algorithm gives

This second approximation has a relative error of less than 1% for angles up to 163.10 degrees. [6]

Approximate formulae for the nonlinear pendulum period

Though the exact period can be determined, for any finite amplitude rad, by evaluating the corresponding complete elliptic integral , where , this is often avoided in applications because it is not possible to express this integral in a closed form in terms of elementary functions. This has made way for research on simple approximate formulae for the increase of the pendulum period with amplitude (useful in introductory physics labs, classical mechanics, electromagnetism, acoustics, electronics, superconductivity, etc. [9] The approximate formulae found by different authors can be classified as follows:

Of course, the increase of with amplitude is more apparent when , as has been observed in many experiments using either a rigid rod or a disc. [12] As accurate timers and sensors are currently available even in introductory physics labs, the experimental errors found in ‘very large-angle’ experiments are already small enough for a comparison with the exact period, and a very good agreement between theory and experiments in which friction is negligible has been found. Since this activity has been encouraged by many instructors, a simple approximate formula for the pendulum period valid for all possible amplitudes, to which experimental data could be compared, was sought. In 2008, Lima derived a weighted-average formula with this characteristic: [9] where , which presents a maximum error of only 0.6% (at ).

Arbitrary-amplitude angular displacement

The Fourier series expansion of is given by [13] [14]

where is the elliptic nome, and the angular frequency.

If one defines can be approximated using the expansion (see OEIS:  A002103 ). Note that for , thus the approximation is applicable even for large amplitudes.

Equivalently, the angle can be given in terms of the Jacobi elliptic function with modulus [15]

For small , , and , so the solution is well-approximated by the solution given in Pendulum (mechanics)#Small-angle approximation.

Examples

The animations below depict the motion of a simple (frictionless) pendulum with increasing amounts of initial displacement of the bob, or equivalently increasing initial velocity. The small graph above each pendulum is the corresponding phase plane diagram; the horizontal axis is displacement and the vertical axis is velocity. With a large enough initial velocity the pendulum does not oscillate back and forth but rotates completely around the pivot.

Compound pendulum

A compound pendulum (or physical pendulum) is one where the rod is not massless, and may have extended size; that is, an arbitrarily shaped rigid body swinging by a pivot . In this case the pendulum's period depends on its moment of inertia around the pivot point.

The equation of torque gives: where: is the angular acceleration. is the torque

The torque is generated by gravity so: where:

Hence, under the small-angle approximation, (or equivalently when ), where is the moment of inertia of the body about the pivot point .

The expression for is of the same form as the conventional simple pendulum and gives a period of [2]

And a frequency of

If the initial angle is taken into consideration (for large amplitudes), then the expression for becomes: and gives a period of: where is the maximum angle of oscillation (with respect to the vertical) and is the complete elliptic integral of the first kind.

An important concept is the equivalent length, , the length of a simple pendulums that has the same angular frequency as the compound pendulum:

Consider the following cases:

Where . Notice these formulae can be particularized into the two previous cases studied before just by considering the mass of the rod or the bob to be zero respectively. Also notice that the formula does not depend on both the mass of the bob and the rod, but actually on their ratio, . An approximation can be made for :

Notice how similar it is to the angular frequency in a spring-mass system with effective mass.

Damped, driven pendulum

The above discussion focuses on a pendulum bob only acted upon by the force of gravity. Suppose a damping force, e.g. air resistance, as well as a sinusoidal driving force acts on the body. This system is a damped, driven oscillator, and is chaotic.

Equation (1) can be written as

(see the Torque derivation of Equation (1) above).

A damping term and forcing term can be added to the right hand side to get

where the damping is assumed to be directly proportional to the angular velocity (this is true for low-speed air resistance, see also Drag (physics)). and are constants defining the amplitude of forcing and the degree of damping respectively. is the angular frequency of the driving oscillations.

Dividing through by :

For a physical pendulum:

This equation exhibits chaotic behaviour. The exact motion of this pendulum can only be found numerically and is highly dependent on initial conditions, e.g. the initial velocity and the starting amplitude. However, the small angle approximation outlined above can still be used under the required conditions to give an approximate analytical solution.

Physical interpretation of the imaginary period

The Jacobian elliptic function that expresses the position of a pendulum as a function of time is a doubly periodic function with a real period and an imaginary period. The real period is, of course, the time it takes the pendulum to go through one full cycle. Paul Appell pointed out a physical interpretation of the imaginary period: [16] if θ0 is the maximum angle of one pendulum and 180° − θ0 is the maximum angle of another, then the real period of each is the magnitude of the imaginary period of the other.

Coupled pendula

Two identical simple pendulums coupled via a spring connecting the bobs. Coupled pendulums.svg
Two identical simple pendulums coupled via a spring connecting the bobs.

Coupled pendulums can affect each other's motion, either through a direction connection (such as a spring connecting the bobs) or through motions in a supporting structure (such as a tabletop). The equations of motion for two identical simple pendulums coupled by a spring connecting the bobs can be obtained using Lagrangian mechanics.

The kinetic energy of the system is: where is the mass of the bobs, is the length of the strings, and , are the angular displacements of the two bobs from equilibrium.

The potential energy of the system is:

where is the gravitational acceleration, and is the spring constant. The displacement of the spring from its equilibrium position assumes the small angle approximation.

The Lagrangian is then which leads to the following set of coupled differential equations:

Adding and subtracting these two equations in turn, and applying the small angle approximation, gives two harmonic oscillator equations in the variables and : with the corresponding solutions where

and , , , are constants of integration.

Expressing the solutions in terms of and alone:

If the bobs are not given an initial push, then the condition requires , which gives (after some rearranging):

See also

Related Research Articles

In classical mechanics, a harmonic oscillator is a system that, when displaced from its equilibrium position, experiences a restoring force F proportional to the displacement x: where k is a positive constant.

<span class="mw-page-title-main">Simple harmonic motion</span> To-and-fro periodic motion in science and engineering

In mechanics and physics, simple harmonic motion is a special type of periodic motion an object experiences by means of a restoring force whose magnitude is directly proportional to the distance of the object from an equilibrium position and acts towards the equilibrium position. It results in an oscillation that is described by a sinusoid which continues indefinitely.

<span class="mw-page-title-main">Spherical coordinate system</span> Coordinates comprising a distance and two angles

In mathematics, a spherical coordinate system is a coordinate system for three-dimensional space where the position of a given point in space is specified by three real numbers: the radial distancer along the radial line connecting the point to the fixed point of origin; the polar angleθ between the radial line and a given polar axis; and the azimuthal angleφ as the angle of rotation of the radial line around the polar axis. (See graphic regarding the "physics convention".) Once the radius is fixed, the three coordinates (r, θ, φ), known as a 3-tuple, provide a coordinate system on a sphere, typically called the spherical polar coordinates. The plane passing through the origin and perpendicular to the polar axis (where the polar angle is a right angle) is called the reference plane (sometimes fundamental plane).

<span class="mw-page-title-main">Trigonometric functions</span> Functions of an angle

In mathematics, the trigonometric functions are real functions which relate an angle of a right-angled triangle to ratios of two side lengths. They are widely used in all sciences that are related to geometry, such as navigation, solid mechanics, celestial mechanics, geodesy, and many others. They are among the simplest periodic functions, and as such are also widely used for studying periodic phenomena through Fourier analysis.

<span class="mw-page-title-main">Solid angle</span> Measure of how large an object appears to an observer at a given point in three-dimensional space

In geometry, a solid angle is a measure of the amount of the field of view from some particular point that a given object covers. That is, it is a measure of how large the object appears to an observer looking from that point. The point from which the object is viewed is called the apex of the solid angle, and the object is said to subtend its solid angle at that point.

<span class="mw-page-title-main">Tautochrone curve</span> Curve for which the time to roll to the end is equal for all starting points

A tautochrone curve or isochrone curve is the curve for which the time taken by an object sliding without friction in uniform gravity to its lowest point is independent of its starting point on the curve. The curve is a cycloid, and the time is equal to π times the square root of the radius over the acceleration of gravity. The tautochrone curve is related to the brachistochrone curve, which is also a cycloid.

In mechanics and geometry, the 3D rotation group, often denoted SO(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition.

<span class="mw-page-title-main">Spherical harmonics</span> Special mathematical functions defined on the surface of a sphere

In mathematics and physical science, spherical harmonics are special functions defined on the surface of a sphere. They are often employed in solving partial differential equations in many scientific fields. The table of spherical harmonics contains a list of common spherical harmonics.

<span class="mw-page-title-main">Inverted pendulum</span> Pendulum with center of mass above pivot

An inverted pendulum is a pendulum that has its center of mass above its pivot point. It is unstable and falls over without additional help. It can be suspended stably in this inverted position by using a control system to monitor the angle of the pole and move the pivot point horizontally back under the center of mass when it starts to fall over, keeping it balanced. The inverted pendulum is a classic problem in dynamics and control theory and is used as a benchmark for testing control strategies. It is often implemented with the pivot point mounted on a cart that can move horizontally under control of an electronic servo system as shown in the photo; this is called a cart and pole apparatus. Most applications limit the pendulum to 1 degree of freedom by affixing the pole to an axis of rotation. Whereas a normal pendulum is stable when hanging downward, an inverted pendulum is inherently unstable, and must be actively balanced in order to remain upright; this can be done either by applying a torque at the pivot point, by moving the pivot point horizontally as part of a feedback system, changing the rate of rotation of a mass mounted on the pendulum on an axis parallel to the pivot axis and thereby generating a net torque on the pendulum, or by oscillating the pivot point vertically. A simple demonstration of moving the pivot point in a feedback system is achieved by balancing an upturned broomstick on the end of one's finger.

<span class="mw-page-title-main">Projectile motion</span> Motion of launched objects due to gravity

Projectile motion is a form of motion experienced by an object or particle that is projected in a gravitational field, such as from Earth's surface, and moves along a curved path under the action of gravity only. In the particular case of projectile motion on Earth, most calculations assume the effects of air resistance are passive.

In quantum physics, the scattering amplitude is the probability amplitude of the outgoing spherical wave relative to the incoming plane wave in a stationary-state scattering process. At large distances from the centrally symmetric scattering center, the plane wave is described by the wavefunction

<span class="mw-page-title-main">Etendue</span> Measure of the "spread" of light in an optical system

Etendue or étendue is a property of light in an optical system, which characterizes how "spread out" the light is in area and angle. It corresponds to the beam parameter product (BPP) in Gaussian beam optics. Other names for etendue include acceptance, throughput, light grasp, light-gathering power, optical extent, and the AΩ product. Throughput and AΩ product are especially used in radiometry and radiative transfer where it is related to the view factor. It is a central concept in nonimaging optics.

<span class="mw-page-title-main">Multiple integral</span> Generalization of definite integrals to functions of multiple variables

In mathematics (specifically multivariable calculus), a multiple integral is a definite integral of a function of several real variables, for instance, f(x, y) or f(x, y, z).

The Kuramoto model, first proposed by Yoshiki Kuramoto, is a mathematical model used in describing synchronization. More specifically, it is a model for the behavior of a large set of coupled oscillators. Its formulation was motivated by the behavior of systems of chemical and biological oscillators, and it has found widespread applications in areas such as neuroscience and oscillating flame dynamics. Kuramoto was quite surprised when the behavior of some physical systems, namely coupled arrays of Josephson junctions, followed his model.

In physics, spherical multipole moments are the coefficients in a series expansion of a potential that varies inversely with the distance R to a source, i.e., as  Examples of such potentials are the electric potential, the magnetic potential and the gravitational potential.

The Wigner D-matrix is a unitary matrix in an irreducible representation of the groups SU(2) and SO(3). It was introduced in 1927 by Eugene Wigner, and plays a fundamental role in the quantum mechanical theory of angular momentum. The complex conjugate of the D-matrix is an eigenfunction of the Hamiltonian of spherical and symmetric rigid rotors. The letter D stands for Darstellung, which means "representation" in German.

<span class="mw-page-title-main">Axis–angle representation</span> Parameterization of a rotation into a unit vector and angle

In mathematics, the axis–angle representation parameterizes a rotation in a three-dimensional Euclidean space by two quantities: a unit vector e indicating the direction of an axis of rotation, and an angle of rotation θ describing the magnitude and sense of the rotation about the axis. Only two numbers, not three, are needed to define the direction of a unit vector e rooted at the origin because the magnitude of e is constrained. For example, the elevation and azimuth angles of e suffice to locate it in any particular Cartesian coordinate frame.

<span class="mw-page-title-main">Rotational diffusion</span> Mechanics concept

Rotational diffusion is the rotational movement which acts upon any object such as particles, molecules, atoms when present in a fluid, by random changes in their orientations. Although the directions and intensities of these changes are statistically random, they do not arise randomly and are instead the result of interactions between particles. One example occurs in colloids, where relatively large insoluble particles are suspended in a greater amount of fluid. The changes in orientation occur from collisions between the particle and the many molecules forming the fluid surrounding the particle, which each transfer kinetic energy to the particle, and as such can be considered random due to the varied speeds and amounts of fluid molecules incident on each individual particle at any given time.

In mathematics, vector spherical harmonics (VSH) are an extension of the scalar spherical harmonics for use with vector fields. The components of the VSH are complex-valued functions expressed in the spherical coordinate basis vectors.

<span class="mw-page-title-main">Elastic pendulum</span>

In physics and mathematics, in the area of dynamical systems, an elastic pendulum is a physical system where a piece of mass is connected to a spring so that the resulting motion contains elements of both a simple pendulum and a one-dimensional spring-mass system. For specific energy values, the system demonstrates all the hallmarks of chaotic behavior and is sensitive to initial conditions.At very low and very high energy, there also appears to be regular motion. The motion of an elastic pendulum is governed by a set of coupled ordinary differential equations.This behavior suggests a complex interplay between energy states and system dynamics.

References

  1. defined by Christiaan Huygens: Huygens, Christian (1673). "Horologium Oscillatorium" (PDF). 17centurymaths. 17thcenturymaths.com. Retrieved 2009-03-01., Part 4, Definition 3, translated July 2007 by Ian Bruce
  2. 1 2 Nave, Carl R. (2006). "Simple pendulum". Hyperphysics. Georgia State Univ. Retrieved 2008-12-10.
  3. Xue, Linwei (2007). "Pendulum Systems". Seeing and Touching Structural Concepts. Civil Engineering Dept., Univ. of Manchester, UK. Retrieved 2008-12-10.
  4. Weisstein, Eric W. (2007). "Simple Pendulum". Eric Weisstein's world of science. Wolfram Research. Retrieved 2009-03-09.
  5. Nelson, Robert; Olsson, M. G. (February 1986). "The pendulum — Rich physics from a simple system". American Journal of Physics. 54 (2): 112–121. Bibcode:1986AmJPh..54..112N. doi:10.1119/1.14703. S2CID   121907349.
  6. 1 2 3 Carvalhaes, Claudio G.; Suppes, Patrick (December 2008), "Approximations for the period of the simple pendulum based on the arithmetic-geometric mean" (PDF), Am. J. Phys. , 76 (12͒): 1150–1154, Bibcode:2008AmJPh..76.1150C, doi:10.1119/1.2968864, ISSN   0002-9505 , retrieved 2013-12-14
  7. Borwein, J.M.; Borwein, P.B. (1987). Pi and the AGM. New York: Wiley. pp. 1–15. ISBN   0-471-83138-7. MR   0877728.
  8. Van Baak, Tom (November 2013). "A New and Wonderful Pendulum Period Equation" (PDF). Horological Science Newsletter. 2013 (5): 22–30.
  9. 1 2 Lima, F. M. S. (2008-09-10). "Simple 'log formulae' for pendulum motion valid for any amplitude". European Journal of Physics. 29 (5): 1091–1098. doi:10.1088/0143-0807/29/5/021. ISSN   0143-0807. S2CID   121743087 via IoP journals.
  10. Lima, F. M. S.; Arun, P. (October 2006). "An accurate formula for the period of a simple pendulum oscillating beyond the small angle regime". American Journal of Physics. 74 (10): 892–895. arXiv: physics/0510206 . Bibcode:2006AmJPh..74..892L. doi:10.1119/1.2215616. ISSN   0002-9505. S2CID   36304104.
  11. Cromer, Alan (February 1995). "Many oscillations of a rigid rod". American Journal of Physics. 63 (2): 112–121. Bibcode:1995AmJPh..63..112C. doi:10.1119/1.17966. ISSN   0002-9505.
  12. Gil, Salvador; Legarreta, Andrés E.; Di Gregorio, Daniel E. (September 2008). "Measuring anharmonicity in a large amplitude pendulum". American Journal of Physics. 76 (9): 843–847. Bibcode:2008AmJPh..76..843G. doi:10.1119/1.2908184. ISSN   0002-9505.
  13. Lawden, Derek F. (1989). Elliptic Functions and Applications. Springer-Verlag. p. 40. ISBN   0-387-96965-9. Eq. 2.7.9:
  14. Reinhardt, W. P.; Walker, P. L. (2010), "Jacobian Elliptic Functions", in Olver, Frank W. J.; Lozier, Daniel M.; Boisvert, Ronald F.; Clark, Charles W. (eds.), NIST Handbook of Mathematical Functions , Cambridge University Press, ISBN   978-0-521-19225-5, MR   2723248 .
  15. "A Complete Solution to the Non-Linear Pendulum". 4 December 2021.
  16. Appell, Paul (July 1878). "Sur une interprétation des valeurs imaginaires du temps en Mécanique" [On an interpretation of imaginary time values in mechanics]. Comptes Rendus Hebdomadaires des Séances de l'Académie des Sciences . 87 (1).

Further reading